Reduced Graphene Oxide Nanocomposite Anodes with

Synthesis and Properties of Si/Ag and Si/Ag/CMS Composites as Anode Materials for Li-ion Batteries. Zha Guojun , Hu Naigen. Silicon 2018 52, ...
0 downloads 0 Views 3MB Size
Subscriber access provided by University of Sydney Library

Article

Si-Mn/Reduced Graphene Oxide Nanocomposite Anodes with Enhanced Capacity and Stability for Lithium-Ion Batteries A Reum Park, Jung Sub Kim, Kwang Su Kim, Kan Zhang, Juhyun Park, Jong Hyeok Park, Joong-Kee Lee, and Pil J. Yoo ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/am404608d • Publication Date (Web): 20 Jan 2014 Downloaded from http://pubs.acs.org on January 27, 2014

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Si-Mn/Reduced Graphene Oxide Nanocomposite Anodes with Enhanced Capacity and Stability for Lithium-Ion Batteries A Reum Park,† Jung Sub Kim,‡ Kwang Su Kim,§ Kan Zhang,† Juhyun Park,║ Jong Hyeok Park,†, § Joong Kee Lee,‡ and Pil J. Yoo†, §,* †

School of Chemical Engineering and §SKKU Advanced Institute of Nanotechnology (SAINT),

Sungkyunkwan University, Suwon 440-746, Republic of Korea. ‡

Advanced Energy materials Processing Laboratory, Center for Energy Convergence Research, Korea

Institute of Science and Technology (KIST), Seoul 130-650, Republic of Korea. ║

School of Chemical Engineering and Materials Science, Chung-Ang University, Seoul 156-756,

Republic of Korea. Email Address: [email protected] RECEIVED DATE (to be automatically inserted after your manuscript is accepted)

ABSTRACT: Although Si is a promising high capacity anode material for Li-ion batteries (LIB), it suffers from capacity fading due to excessively large volumetric changes upon Li insertion. Nanocarbon materials have been used to enhance the cyclic stability of LIB anodes, but they have an inherently low specific capacity. To address these issues, we present a novel ternary nanocomposite of Si, Mn, and reduced graphene oxide (rGO) for LIB anodes, in which Si-Mn alloy offers high capacity characteristics and embedded rGO nanosheets confers structural stability.

Si-Mn/rGO ternary

nanocomposites were synthesized by mechanical complexation and subsequent thermal reduction of

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 19

mixtures of Si nanoparticles, MnO2 nanorods, and rGO nanosheets. Resulting ternary nanocomposite anodes displayed a specific capacity of 600 mAh/g with ~ 90 % capacity retention after 50 cycles at a current density of 100 mA/g. The enhanced performances are attributed to facilitated Li ion reactions with the MnSi alloy phase and the formation of a structurally-reinforced electroconductive matrix of rGO nanosheets. The ternary nanocomposite design paradigm presented in this study can be exploited to develop high capacity and long-life anode materials for versatile LIB applications.

KEYWORDS: Silicon, Manganese, Reduced graphene oxide, Lithium-ion batteries, Anodes

INTRODUCTION The increasing demand for high-performance portable electronic devices, electric vehicles, and energy integration systems, has spurred efforts to develop next-generation lithium ion batteries (LIB) with greater capacity and energy density.1 To meet this requirement, Si is the material of choice for node materials due to its high theoretical capacity (4,200 mAh/g), which is about 10 times greater than that of conventional graphite anodes.2 However, Si-based anode materials suffer from large volume changes (~ 300 %) and the subsequent accumulation of excessive stress during lithiation/delithiation cycles, resulting in structural collapse of the electrode and drastic capacity fading.3 To circumvent this issue, Si alloys with a Li-inactive element, such as NiSi, MgSi, or FeSi, have been employed to decrease structural changes inside the anodes.4-6 The alloyed transition metal species mitigates Si pulverization in response to repeated cycles, and provides mechanical support for the electrochemical activity of the electrodes. However, although the cycling characteristics of Si-alloyed anodes are substantially better than those of regular Si anodes, incomplete interconnections between the alloyed domains retard cell performance, thus, capacity fading nevertheless occurs.

Therefore, graphite is generally added as an

electroconductive matrix to promote electrical interconnections between active materials.7,8 To achieve uniform integration of graphite into Si alloys, a high temperature ball milling process is

ACS Paragon Plus Environment

2

Page 3 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

usually required.9,10 This harsh process results in the formation of silicon oxides or other Li-inactive compounds, that can undermine electrode performance. In addition, graphite may have limited capacity due to a relatively large irreversible capacity loss in the first cycle, which varies depending on the internal porosity or effective surface area.11 For these reasons, sheet-like structured graphene has recently received much attention because of its excellent specific capacity and cyclic reversibility afforded by extra spaces for further Li-ion intercalation availability within the stacked graphene layers.12 In addition, very large theoretical surface area (2,630 m2/g), outstanding electrical conductivity, and mechanical rigidity of graphene make it an ideal material to impart conductivity and structural stability to Si-based anode materials.13-16 Meanwhile, manganese, which is a less toxic and naturally abundant transition metal species, can be also used for LIB anode applications. Though it is Li-inactive in itself, particularly when hybridized with Si, Mn can form a Li-active Si-Mn alloy phase that is structurally stable during lithiation/delithiation cycles.17,18 Therefore, in this study, we used Si-Mn alloy and graphene nanosheets for LIB anodes to simultaneously obtain enhanced specific capacity and cyclic stability.

Si-

Mn/graphene nanocomplexes were prepared by mild mechanical ball-milling of Si nanoparticles, MnO2 nanorods, and reduced graphene oxide (rGO) nanosheets, followed by thermal treatment to reduce MnO2 (Mn4+) to Mn (Mn0). The formation of the MnSi phase within Si-Mn alloys results in high specific capacity for electrode applications. Furthermore, uniform interdigitation of rGO nanosheets inside the MnSi alloy phase efficiently suppresses self-aggregation or electrical disconnection problems within the Li-active domains. This structural manipulation enhances the electrochemical activity of electrodes without sacrificing the structural framework, thereby resulting in enhanced cyclic stability.

EXPERIMENTAL SECTION Synthesis of manganese oxide nanorods MnO2 nanorods were synthesized via a quick-precipitation procedure.19 Manganese(II) chloride tetrahydrate (MnCl2 • 4H2O) (0.27 g, Aldrich) was dispersed in

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 19

isopropyl alcohol (50 mL, Samchun Chemical, Korea) in a round-bottomed flask equipped with a reflux condenser. The solution was heated to 83°C with vigorous stirring, and KMnO4 (0.15 g, Samchun Chemical) dissolved in 5 mL of deionized (DI) water was then added rapidly to the above boiling solution. Immediately, a large amount of black precipitate was produced. After refluxing for 0.5 h, the mixture was allowed to cool to room temperature. After washing the precipitate with DI water and ethanol several times, followed by drying at 60°C, MnO2 nanorods were finally obtained. Synthesis of graphene oxide (GO) and reduced graphene oxide (rGO) Graphene oxide (GO) was synthesized from graphite powder (< 20 µm, synthetic, Aldrich) using the modified Hummers method.20,21 First, 0.5 g of graphite flakes and 0.5 g of NaNO3 (99.0 %, Yakuri Pure Chemicals, Japan) were stirred into 24 ml of H2SO4 (95.0 %, Samchun Chemical). Then, 3g of KMnO4 (99.3%, Samchun Chemical) was added slowly while preventing the temperature of the suspension from exceeding 20°C in ice bath. After stirring continuously the mixture for 1 h at 35°C, 40 mL of DI water was slowly added to dilute mixture and the temperature was raised to 90°C. To reduce the residual permanganate and manganese dioxide to colorless soluble manganese sulfate, 5 mL of H2O2 (34.5 %, Samchun Chemical) was added and the suspension was filtered with 2 L of DI water. The obtained a yellowbrown suspension was exfoliated to produce single layer graphite oxide using ultrasonication and the unexfoliated precipitation was removed by centrifugation.

The resulting aqueous suspension was

completely dried at 60°C, followed by thermal reduction in a tube furnace at 800°C for 1 h under a H2/Ar (1:3) atmosphere, yielding a powder of rGO nanosheets. Preparation of Si-Mn/rGO ternary nanocomposites Si-Mn nanocomposites were prepared by mechanical ball milling of Si nanoparticles (< 100 nm, Aldrich) and MnO2 nanorods at the mass ratio of 1:1. Ball milling was performed in a planetary ball-milling machine (8000D Dual Mixer/Mill, SPEX SamplePrep, NJ) for the relatively short time of 30 min. To prepare ternary nanocomposites, the prepared rGO was added to the Si-Mn nanocomposites (Si-Mn : rGO = 80 : 20 in wt%) and the mixture was ball-milled for 20 min. The resulting ternary composite was annealed at 800°C for 3 h under an Ar

ACS Paragon Plus Environment

4

Page 5 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

atmosphere to reduce MnO2. For a comparison, Si-Mn binary composite without including rGO was also prepared by ball-milling and subsequent thermal annealing process. Structural characterization Morphologies of Si-Mn and Si-Mn/rGO nanocomposite particles were characterized by field-emission scanning electron microscopy (FESEM, JSM-7600F, JEOL, Japan), and high-resolution transmission electron microscopy (HRTEM, JEM-3010, JEOL, Japan). Powder X-ray diffraction (XRD) patterns were obtained (D8 Focus, Bruker AXS, Germany) with Cu Kα radiation (λ = 1.5406 Å) in the 2θ range from 3 to 80° with a step size of 0.02° s−1. Raman spectra were collected using a micro-Raman spectrometer system (ALPHA 300M, WITec, Germany). Characterization of electrochemical properties Electrochemical experiments were performed using CR2032 coin cells.

To prepare electrodes, Si-Mn/rGO ternary nanocomposites were mixed with

conductive black (Super-P) and poly(acrylic acid) (PAA, Mw ~ 250,000, 35 wt.% in H2O, Aldrich) in ethanol at a mass ratio of 70:15:15 into a homogeneous slurry. The mixed slurry was uniformly pasted on Cu foil to have ~1.5 mg/cm2 of an active mass with a film thickness at 36 µm. Coin cells were then assembled with Li metal as the counter electrode and polypropylene (PP, Celgard 2400, Celgard, NC) film as the separator. 1 M LiPF6 solution in a 1:1:1 (v/v %) mixture of ethylene carbonate (EC), ethylmethyl carbonate (EMC), and diemethyl carbonate (DMC) was used as the electrolyte. Electrochemical

characteristics

of

the

batteries

were

measured

with

a

multi-channel

potentiostat/galvanostat (WMPG 1000, WonATech, Korea) at room temperature. Charge–discharge cycles were characterized galvanostatically in the potential range of 0.01 – 3.25 V (vs. Li/Li+) at current densities ranging from 100 − 2,000 mA/g. Cyclic voltammetry was carried out in the potential range of 0.01–3.25 V at a scan rate of 0.1 mA/ s. Structural change of ternary nanocomposites after chargingdischarging cycles was observed with SEM. Coin cells were disassembled in a glove box and the separated electrode film was immersed in DMC for 10 min to remove LiPF6 electrolyte, followed by complete drying for SEM observation.

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 19

RESULTS AND DISCUSSION

Figure 1. Schematic procedure of fabricating the Si-Mn/rGO ternary nanocomposites

Synthesis procedure used to produce Si-Mn/rGO ternary nanocomposites is shown schematically in Fig. 1. First, Si nanoparticles (< 100 nm in diameter) were mechanically mixed with MnO2 nanorods (20 − 50 nm in diameter, aspect ratio ~ 10) to form a homogeneously complexed phase. We used onedimensionally structured MnO2 nanorods to maximize interfacial contact with Si nanoparticles during complexation to suppress the self-aggregation that is frequently encountered in mixing of spherical nanoparticles.22,23 Then, dried rGO nanosheet powder was added to the pretreated sample followed by mixing to generate a ternary nanocomposite. Graphene oxide (GO) nanosheets were thermally reduced to form rGO prior to mixing. This procedure prevents oxidation of Si nanoparticles, moisture uptake and subsequent gel-phase formation, which are all undesirable for attaining uniform complexation.24 Then, to reduce MnO2 to Mn in the presence of carbon (rGO) and to promote formation of a stable MnSi alloy phase, samples were thermally annealed at 800°C for 3 h in an Ar-purged environment (byproduct gas of CO2 was thereby readily removed25). This process yielded Si-Mn/rGO ternary nanocomposites with a Li-active MnSi alloy phase encompassed by rGO nanosheets.

ACS Paragon Plus Environment

6

Page 7 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 2. Comparative SEM images of (a, b) Si-Mn binary nanocomposites. (c, d) Si-Mn/rGO ternary nanocomposites.

Morphological characteristics of Si-Mn/rGO nanocomposites were observed by scanning electron microscopy (SEM). As shown in Figs. 2a and 2b, when Si nanoparticles and MnO2 nanorods were complexed by ball milling and thermally annealed in the absence of rGO nanosheets (Si-Mn binary nanocomposites), several micrometer-sized particles were generated, and Si-Mn alloys as a minor phase were incorporated inside the Si matrix. Due to the homogeneous mixing provided by ball milling, particles were densely agglomerated and the surface was relatively smooth. In contrast, when rGO nanosheets were added (Si-Mn/rGO ternary nanocomposites), the reduction of MnO2 to Mn was greatly facilitated and a highly textured surface was obtained, as shown in Figs. 2c and 2d, indicating successful embedment of rGO nanosheets inside the Si-Mn alloy matrix. Here, it should be noted that an excess amount of Si was mixed with MnO2 (atomic ratio between Si and Mn ~ 3) to ensure full conversion of Mn species to Si-Mn alloy phase. More importantly, this condition ensured dominant formation of the MnSi phase rather than other Li-inactive alloys.26 Because uniform complexation of rGO nanosheets simultaneously increases electrical interconnections between Li-active domains, we expected the ternary ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 19

nanocomposites to show enhanced electrochemical performance in the context of LIB anode applications.

Figure 3. Crystallographic and chemical structure analysis of Si-Mn/rGO ternary nanocomposites. (a) X-ray diffractograms for complexed product of Si/MnO2/rGO, Si/MnO2, and MnO2/rGO, respectively, and for individual rGO and Si (from top to bottom). (b) Raman spectra for complexed product of Si/MnO2/rGO and for individual Si, reduced MnO2 (Mn), and rGO (from top to bottom).

As-prepared samples were characterized by X-ray diffraction (XRD) measurements to obtain crystallographic information (Fig. 3a).

For the complexation product of Si/MnO2 (Si-Mn binary

ACS Paragon Plus Environment

8

Page 9 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

nanocomposite), peaks at 27.7°, 34.0°, 39.5°, 44.4°, and 48.9° correspond to the (110), (111), (200), (210), and (211) planes of the MnSi alloy phase, respectively, according to JCPDS 65-3297. The Si-Mn binary nanocomposite was also characterized by peaks at 28.4°, 47.6°, and 56.1° that were indexed to (111), (220), and (311) planes of Si crystals, confirming the presence of residual Si in the matrix.27 These peaks were also clearly observed in the Si-Mn/rGO ternary nanocomposites for the complexation of Si/MnO2/rGO, whereas the peak associated with rGO was hard to distinguish because of its less ordered status and correspondingly weaker intensity than other peaks. Notably, peaks associated with Mn2SiO4 phase formation observed in the Si-Mn binary composite were almost vanished in the ternary system, indicating a role of rGO as a reducing agent. This could be more clearly confirmed by the result of complexation between rGO and MnO2, in which the partial reduction of MnO2 to MnO took place. These observations indicated that the rGO complexation successfully facilitated the reduction of MnO2 in the presence of Si to form the MnSi alloy phase, and at the same time, contributed to promote electrical conductivity by creating internal interconnections. In addition to these peaks, very minor peaks corresponding to Mn5Si3 and Mn11Si19 from Si-Mn alloy samples were also observed.28,29 However, the peaks with the highest intensities were mostly contributed by the MnSi alloy phase. These results together with the SEM images shown in Fig. 2 displayed that the Si-Mn alloy consisted of micro-scale particulates of Si matrix and uniformly embedded nanoparticles with smaller domains of Liactive MnSi phase. The chemical structure of the Si-Mn/rGO nanocomposites was elucidated by Raman spectra analysis. Consistent with the XRD results, Raman spectra for the as-prepared samples shown in Fig. 3b comprised characteristic Raman peaks for individual species. Here, the spectrum for Mn was obtained from the thermally reduced MnO2 under a H2/Ag atmosphere. Peaks observed at 517 and 651 cm−1 were indexed to Si and Mn, respectively, implying the successful formation of Si-Mn alloy.30-32 Since Si was complexed in excess and occupied the major matrix of ternary nanocomposites, the Si peak was more pronounced than the Mn peak. In the case of rGO, both D (1,350 cm−1) and G (1,580 cm−1) bands,

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 19

which are generally associated with in-plane sp3 defects and sp2 domains in carbon, respectively, were observed.33 These characteristic peaks of rGO were also observed in the ternary nanocomposites, though the captured intensities were quite weak. The intensity ratio of the D band to the G band (ID/IG), which is used as an indicator of graphene disorder, decreased slightly from 1.08 in pure rGO to 1.00 in ternary nanocomposites, presumably indicating enlargement of the graphitic domain during complexation with the Si-Mn phase.

Figure 4. (a) TEM observation of the Si-Mn/rGO ternary nanocomposites. (b) HRTEM image and FFT analysis (inset). (c) TEM image of the Si-Mn/rGO was selected for EDX mappings of (d) Si, (e) Mn, and (f) C.

To characterize the internal structure of Si-Mn/rGO ternary nanocomposites more precisely, we investigated the samples using high resolution-TEM (HRTEM). As shown in Fig. 4a, Si-Mn alloy was ACS Paragon Plus Environment

10

Page 11 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

surrounded by multilayered rGO nanosheets, consistent with the results from SEM observations. Furthermore, the highly magnified image presented in Fig. 4b revealed a lattice spacing of 0.20 nm that corresponds to the (210) plane of MnSi, verifying that the structural characteristics of the Si-Mn alloy remained intact despite rGO complexation. The inset Fast Fourier Transform (FFT) analysis supported the crystallographic information of MnSi in the Si-Mn/rGO ternary nanocomposites, in which (110), (111), (200), and (210) planes of MnSi were indexed. We further investigated the complexation uniformity of Si-Mn/rGO nanocomposites using energy dispersive X-ray spectroscopy (EDX). As shown in Fig. 4c, black-colored domains that corresponded to the MnSi phase were randomly embedded inside the matrix of Si (expressed as the gray-colored background). The light gray-colored regions in Fig. 4c indicate the outermost layer of the surrounding rGO. EDX results for Si, Mn, and C presented in Figs. 4d − 4f demonstrated uniform incorporation of the MnSi alloy phase inside the rGO nanosheetwrapped Si matrix. Electrochemical performances of the Si-Mn/rGO nanocomposites as LIB anodes were tested; results are presented in Fig. 5. First, major reaction regions were investigated using cyclic voltammetry (CV) measurements as shown in Fig. 5a. Presented curves were obtained from the first to tenth cycles for ternary nanocomposites in the voltage range of 0.01 − 3.25 V (versus Li+/Li) at a scanning rate of 0.1 mV/s.

Because current density and charge transfer increased gradually during the early

charge/discharge cycles, redox peaks in the first cycle were weak and occupied a broad region.34 Cathodic peaks at 0.45 and 1.0 V indicated the formation of lithiated MnSi (LixMnSi). Other cathodic peaks at 0.051 and 0.20 V were assigned to the Li-Si alloying (LiySi) reaction. In terms of anodic reactions, delithiation of the LixMnSi phase as associated with a slight peak at 1.3 V was observed after the second cycle.18 Discrete anodic peaks at 0.31 and 0.50 V corresponded to the Li-Si dealloying process.35,36 These characteristics matched well the results from microscopic observation of SEM and TEM, wherein MnSi alloy particles were embedded inside the Si matrix. Though very minor inclusion of other Li-active alloy phases such as Mn5Si3 or Mn11Si19 were accompanied during composite

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 19

formation,37 overall electrode performance was dominantly determined by the MnSi phase of the major species in the ternary composites.

Figure 5. Electrochemical performances of Si-Mn/rGO ternary nanocomposite anodes. (a) Cyclic voltammetry. (b) Galvanostatic charging-discharging profiles in the potential range of 0.01 – 3.25 V (vs. Li/Li+) at current density of 100 mA/g. (c) Capacity retention and coulombic efficiency of ternary nanocomposite electrodes during 50 cycles at 100 mA/g. (d) Galvanostatic charging-discharging profiles of the first cycle with varying current density. Voltage profiles were obtained after stabilizing the electrodes with applying 10 cycles at 100 mA/g. (e) Rate capabilities of ternary nanocomposite electrodes.

ACS Paragon Plus Environment

12

Page 13 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Next, galvanostatic charge-discharge profiles were investigated as shown in Fig. 5b. Capacitancevoltage measurements were performed at a current density of 100 mA/g in the active electrode within a cut-off voltage window of 0.01 − 3.25 V versus Li+/Li in 1.0 M LiPF6. Specific capacity values were estimated based on the net weight of active materials (70 wt% of the total electrode mass). In the first charging cycle, we observed an extended slope between 0.01 and 0.1 V associated with the formation of amorphous LiySi. In the second cycle, a voltage plateau of 0.6 V was observed, corresponding to lithiation reaction to form LixMnSi species. Complementarily, in the discharging process, delithiation process of LixMnSi was observed in the voltage range between 1.0 and 2.0 V.18 In the first cycle, the charge and discharge capacity were measured as 1,033 and 668 mAh/g, respectively, and coulombic efficiency reached 64.6%. This large irreversible loss in the first cycle implied the solid electrolyte interface (SEI) formation between the electrode and electrolyte.38 After 50 cycles, however, these values had stabilized to 609 and 595 mAh/g with a coulombic efficiency of 97.6%, indicating good reversible characteristics. The obtained result was greater than previously reported values for general composites of Si/Mn/C, wherein initial discharge capacity of 463 mAh/g decreased to 387 mAh/g after 40 cycles.26 This difference could be ascribed to a formation of Li-active MnSi alloy phase in our ternary nanocomposites. In addition, capacity retention of Si-Mn/rGO nanocomposites was observed to be substantially good, as shown in Fig. 5c. We attributed this pronounced cyclic stability to the formation of MnSi alloy phase and uniform embedment of electroconductive rGO nanosheets. As a result, 89.1% of the initial discharge capacity was retained after 50 cycle operations. Furthermore, the coulombic efficiency showed good reversibility (> ~ 98 %) after around 10 cycles. It should be noted that a slight increase in the capacity during initial few cycles could be ascribed to the activation process of Si matrix in the electrode. The increased intensity of redox peaks with increasing cycles as shown in CV measurements also supported this observation.30 In addition, as revealed in Fig. 5c, capacity retention characteristics

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 19

of Si-Mn/rGO ternary and Si-Mn binary nanocomposites exhibited similar trend, which indirectly proved cyclic stability of the MnSi alloy phase. The enhanced electrical interconnection and facilitated electron transport in the ternary nanocomposites due to the rGO embedment further increased the specific capacity while retaining the cyclic stability. The structural stability of the ternary nanocomposites was also demonstrated by the excellent rate capability characteristics shown in Fig. 5d and 5e. Here, charge-discharge cycles were measured while varying the current density between 100 and 2,000 mA/g. Notably, for the case of Si-Mn/rGO ternary nanocomposites, discharge capacity rarely changed despite an increase in the current density, and coulombic efficiency was consistently maintained at between 97 and 98%. The structural stability of SiMn/rGO nanocomposites was also investigated with SEM observation for tested cells. As shown in Figure 6, internal morphologies of the electrode showed little difference between samples after 1 cycle and 30 cycles of operation, respectively.

Figure 6. Comparative SEM observation of the internal structure of Si-Mn/rGO ternary nanocomposite electrodes. (a) After 1 cycle. (b) After 30 cycles.

ACS Paragon Plus Environment

14

Page 15 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

CONCLUSIONS To summarize, we developed a novel ternary nanocomposite composed of Si, Mn, and reduced graphene oxide that showed highly enhanced performance as a Li-ion battery anode. Preferential formation of Li-active MnSi alloy phase on the electroconductive rGO matrix ensured high capacity because it facilitated electron transport through interconnected rGOs.

Furthermore, uniformly-

embedded electroconductive rGO nanosheets inside the ternary nanocomposites simultaneously promoted specific capacity and cyclic stability.

As a result, Si-Mn/rGO ternary nanocomposites

exhibited a stabilized specific capacity greater than 600 mAh/g after 50 cycles of operation while retaining about 90% of their initial capacity. We anticipate that further structural control of ternary nanocomposites by creating an ordered structure would further enhance the electrochemical performance of these nanocomposites as high performance anode materials for Li-ion battery applications.

ACKNOWLEDGEMENTS This work was supported by research grants of NRF (2012M1A2A2671795, 2012S1A2A1A01031215), Global Frontier R&D Program on Center for Multiscale Energy System (2012M3A6A7055540), and Basic Science Research Program (2012-0009158) funded by the National Research Foundation under the Ministry of Science, ICT & Future, Korea. The authors also acknowledge the funding support from LG Yonam Foundation.

REFERENCES (1) Tarascon, J. -M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. (2) Obrovac, M. N.; Christensen, L. Structural Changes in Silicon Anodes during Lithium Insertion/Extraction. Electrochem. Solid-State Lett. 2004, 7, A93–A96. (3) Yang, J.; Wang, B. F.; Wang, K.; Liu, Y.; Xie, J. Y.; Wen, Z. S. Si/C Composites for High Capacity ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 19

Lithium Storage Materials. Electrochem. Solid-State Lett. 2003, 6, A154–A156. (4) Wang, G. X.; Sun, L.; Bradhurst, D. H.; Zhong, S.; Dou, S. X.; Liu, H. K. Nanocrystalline NiSi alloy as an anode material for lithium-ion batteries. J. Alloys Compd. 2000, 306, 249–252. (5) Roberts, G. A.; Cairns, E. J.; Reimer, J. A. Mechanism of Lithium Insertion into Magnesium Silicide. J. Electrochem. Soc. 2004, 151, A493–A496. (6) Wang, G. X.; Sun, L.; Bradhurst, D. H.; Zhong, S.; Dou, S. X.; Liu, H. K. Innovative nanosize lithium storage alloys with silica as active centre. J. Power Sources 2000, 88, 278–281. (7) Wang, X.; Wen, Z.; Liu, Y.; Xu, X.; Lin, J. Preparation and characterization of a new nanosized silicon-nickel-graphite composite as anode material for lithium ion batteries . J. Power Sources 2009, 189, 121–126. (8) Lee, H. -Y.; Lee, S. -M. Graphite-FeSi alloy composites as anode materials for rechargeable lithium batteries. J. Power Sources 2002, 112, 649–654. (9) Lee, H. -Y.; Lee, S. -M. Carbon-coated nano-Si dispersed oxides/graphite composites as anode material for lithium ion batteries. Electrochem. commun. 2004, 6, 465–469. (10) Feng, X.; Yang, J.; Gao, P.; Wang, J.; Nuli, Y. Facile approach to an advanced nanoporous silicon/carbon composite anode material for lithium ion batteries. RSC Adv. 2012, 2, 5701–5706. (11) Shim, J.; Striebel, K. A. The dependence of natural graphite anode performance on electrode density. J. Power Sources 2004, 130, 247–253. (12) Yoo, E.; Kim, J.; Hosono, E.; Zhou, H. -s.; Kudo, T.; Honma, I. Large Reversible Li Storage of Graphene Nanosheet Families for Use in Rechargeable Lithium Ion Batteries. Nano Lett. 2008, 8, 2277–2282. (13) Chou, S. -L.; Wang, J. -Z.; Choucair, M.; Liu, H. -K.; Stride, J. A.; Dou, S. -X. Enhanced reversible lithium storage in a nanosize silicon/graphene composite. Electrochem. commun. 2010, 12, 303–306. (14) Zhou, X.; Yin, Y. -X.; Cao, A. -M.; Wan, L. -J.; Guo, Y. -G. Efficient 3D Conducting Networks Built by Graphene Sheets and Carbon Nanoparticles for High-Performance Silicon Anode. ACS Appl. Mater. Interfaces 2012, 4, 2824–2828. (15) Chang, J.; Huang, X.; Zhou, G.; Cui, S.; Hallac, P. B.; Jiang, J.; Hurley, P. T.; Chen, J. Adv. Mater. 2013, DOI: 10.1002/adma.201302757. (16) Wang, B.; Li, X.; Zhang, X.; Luo, B.; Jin, M.; Liang, M.; Dayeh, S. A; Picraux, S. T.; Zhi, L. Adaptable Silicon-Carbon Nanocables Sandwiched between Reduced Graphene Oxide Sheets as ACS Paragon Plus Environment

16

Page 17 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Lithium Ion Battery Anodes. ACS Nano 2013, 7, 1437–1445. (17) Park, J. H.; Jeong, D. H.; Cha, S. M.; Sun, Y. -K.; Yoon, C. S. Electrochemical behaviour of Heusler alloy Co2MnSi for secondary lithium batteries. J. Power Sources 2009, 188, 281–285. (18) Vediappan, K.; Lee, C. W. Synthesis and Electrochemical Characterization of Si-Mn Alloy Anode Materials for High Energy Lithium Secondary Batteries. J. Nanosci. Nanotechnol. 2011, 11, 5969–5974. (19) Chen, S.; Zhu, J.; Han, Q.; Zheng, Z.; Yang, Y.; Wang, X. Shape-Controlled Synthesis of OneDimensional MnO2 via a Facile Quick-Precipitation Procedure and its Electrochemical Properties. Cryst. Growth Des. 2009, 9, 4356–4361. (20) Park, J. S.; Cho, S. M.; Kim, W. -J.; Park, J.; Yoo, P. J. Fabrication of Graphene Thin Films Based on Layer-by-Layer Self-Assembly of Functionalized Graphene Nanosheets. ACS Appl. Mater. Interfaces 2011, 3, 360–368. (21) Sher Shah, Md. S. A.; Park, A. R.; Zhang, K.; Park, J. H.; Yoo, P. J. Green Synthesis of Biphasic TiO2-Reduced Graphene Oxide Nanocomposites with Highly Enhanced Photocatalytic Activity. ACS Appl. Mater. Interfaces 2012, 4, 3893–3901. (22) Wang, X.; Wen, Z.; Liu, Y. A novel nanosized silicon-based composite as anode material for high performance lithium ion batteries. Electrochim. Acta 2011, 56, 1512–1517. (23) Szczech, J. R.; Jin, S. Nanostructured silicon for high capacity lithium battery anodes. Energy Environ. Sci. 2011, 4, 56–72. (24) Xiang, H.; Zhang, K.; Ji, G.; Lee, J. Y.; Zou, C.; Chen, X.; Wu, J. Graphene/nanosized silicon composites for lithium battery anodes with improved cycling stability. Carbon 2011, 49, 1787– 1796. (25) Welham, N. J. Activation of the carbothermic reduction of manganese ore. Int. J. Miner. Process. 2002, 67, 187–198. (26) Zuo, P.; Yin, G.; Zhao, J.; Ma, Y.; Cheng, X.; Shi, P.; Takamura, T. Electrochemical reaction of the SiMn/C composite for anode in lithium ion batteries. Electrochim. Acta 2006, 52, 1527–1531. (27) Lee, J. K.; Smith, K. B.; Hayner, C. M.; Kung, H. H. Silicon nanoparticles–graphene paper composites for Li ion battery anodes. Chem. Commun. 2010, 46, 2025–2027. (28) Higgins, J. M.; Ding, R.; Jin, S. Synthesis and Characterization of Manganese-Rich Silicide (αMn5Si3, β-Mn5Si3, and β-Mn3Si) Nanowires. Chem. Mater. 2011, 23, 3848–3853. (29) Udono, H.; Nakamori, K.; Takahashi, Y.; Ujiie, Y.; Ohsugi, I. J.; Iida, T. Solution Growth and ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 19

Thermoelectric Properties of Single-Phase MnSi1.75-x. J. Electron. Mater. 2011, 40, 1165–1170. (30) Bernard, M. -C.; Goff, A. H. -L.; Thi, B. V.; De Torresi, S. C. Electrochromic Reactions in Manganese Oxides. J. Electrochem. Soc. 1993, 140, 3065–3070. (31) Buciuman, F.; Patcas, F.; Craciun, R.; Zahn, D. R. T. Vibrational spectroscopy of bulk and supported manganese oxides. Phys. Chem. Chem. Phys. 1999, 1, 185–190. (32) Tao, H. -C.; Fan, L. -Z.; Mei, Y.; Qu, X. Self-supporting Si/Reduced Graphene Oxide nanocomposite films as anode for lithium ion batteries. Electrochem. commun. 2011, 13, 1332–1335. (33) Ferrari, A. C. Raman spectroscopy of graphene and graphite: Disorder, electron-phonon coupling, doping and nonadiabatic effects. Solid State Commun. 2007, 143, 47–57. (34) Shin, H. -C.; Corno, J. A.; Gole, J. L.; Liu, M. Porous silicon negative electrodes for rechargeable lithium batteries. J. Power Sources 2005, 139, 314–320. (35) Yao, Y.; McDowell, M. T.; Ryu, I.; Wu, H.; Liu, N.; Hu, L.; Nix, W. D.; Cui, Y. Interconnected Silicon Hollow Nanospheres for Lithium-Ion Battery Anodes with Long Cycle Life. Nano Lett. 2011, 11, 2949–2954. (36) Ji, J.; Ji, H.; Zhang, L. L.; Zhao, X.; Bai, X.; Fan, X.; Zhang, F.; Ruoff, R. S. GrapheneEncapsulated Si on Ultrathin-Graphite Foam as Anode for High Capacity Lithium-Ion Batteries. Adv. Mater. 2013, 25, 4673–4677. (37) Anani, A.; Huggins, R. A. Multinary alloy electrodes for solid state batteries Ⅰ. A phase diagram approach for the selection and storage properties determination of candidate electrode materials. J. Power Sources 1992, 38, 351–362. (38) Zhou, M.; Cai, T.; Pu, F.; Chen, H.; Wang, Z.; Zhang, H.; Guan, S. Graphene/Carbon-Coated Si Nanoparticle Hybrids as High-Performance Anode Materials for Li-Ion Batteries. ACS Appl. Mater. Interfaces 2013, 5, 3449–3455

ACS Paragon Plus Environment

18

Page 19 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

SYNOPSIS TOC

Si-Mn/Reduced Graphene Oxide Nanocomposite Anodes with Enhanced Capacity and Stability for Lithium-Ion Batteries A Reum Park, Jung Sub Kim, Kwang Su Kim, Kan Zhang, Juhyun Park, Jong Hyeok Park, Joong Kee Lee, and Pil J. Yoo

ACS Paragon Plus Environment

19